首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Sono-dispersion of Ni, Co and Ni–Co over Al2O3–MgO with Al/Mg ratio of 1.5 was prepared and tested for dry reforming of methane. The samples were characterized by XRD, FESEM, PSD, EDX, TEM, BET and FTIR analyses. In order to assess the effect of ultrasound irradiation, Ni–Co/Al2O3–MgO with Co content of 8% prepared via sonochemistry and impregnation methods. The sono-synthesized sample showed better textural properties and higher activity than that of impregnated one. Comparison of XRD patterns indicated that the NiO peaks became broader by increasing Co content over the support. The FESEM images displayed the particles are small and well-dispersed as a result of sonochemistry method. Also, EDX analysis demonstrated better dispersion of Ni and Co as a result of sonochemistry method in confirmation of XRD analysis. The sono-synthesized Ni–Co/Al2O3–MgO as a superior nanocatalyst with Co content of 3% illustrates much higher conversions (97.5% and 99% for CH4 and CO2 at 850 °C), yields (94% and 96% for H2 and CO at 850 °C) and 0.97 of H2/CO molar ratio in all samples using an equimolar feed ratio at 850 °C. During the 1200 min stability test, H2/CO molar ratio remained constant for the superior nanocatalyst.  相似文献   

2.
Polycrystalline samples of M3(BO3)F3 (M=Fe, Co, Ni), isostructural with nocerite Mg3(BO3)(OH,F)3, have been prepared in supercritical hydrothermal conditions. These compounds represent with boracites, M3B7O13F (M=Mg, Cr, Mn, Fe, Co, Zn), the only transition metal fluoride borates known to date. Co3(BO3)F3 and Ni3(BO3)F3 are antiferromagnetic with TN=17(2) and 40(2) K, respectively. Spin-flop transitions at BC1=4.0 T and BC2=7.5 T occur at 1.6 K in Co3(BO3)F3, while a parasitic ferromagnetism (0.02 μB/Ni2+ at 1.6 K) appears below TN in Ni3(BO3)F3. The magnetic structures consist of three spin sub-lattices of double rutile-type ferromagnetic chains.  相似文献   

3.
Magnetic and magnetocaloric properties of polycrystalline samples of Gd3Co and Gd3Ni have been studied. Both these compounds are antiferromagnets and undergo metamagnetic transitions in the antiferromagnetic phase. The Neel temperatures are found to be 128 and 99 K, respectively for Gd3Co and Gd3Ni. Though both these compounds have the same crystal structure, their magnetic structures seem to be different. It is found that Gd3Ni possesses larger magnetic anisotropy compared to Gd3Co. The maximum values of isothermal magnetic entropy change are 11 and 18.5 J Kg−1 K−1 for Gd3Co and Gd3Ni respectively, while their refrigerant capacities are 2.6 J cm−3 for both of them. Magnetic entropy change in the paramagnetic region shows a quadratic dependence on the magnetic field in the case of Gd3Ni, indicating the presence of spin fluctuations above the Neel temperature.  相似文献   

4.
Single-phase broad-band red-emitting Ca3Si2O7:Eu2+ phosphors, with photoluminescence features that qualify them as candidates for white light-emitting diodes applications, were successfully synthesized via a modified solid-state reaction method that employed H3BO3 as a flux. The phosphors produced have an intense broad red emission band, with a peak at 603 nm, a full width at half maximum of 110 nm, and color coordinates of (0.550, 0.438). Concentration quenching occurred at 0.01 mol Eu2+. The discussion of the results shows that Eu2+ ions should be accommodated at the Ca-sites of the lattice, dipole–dipole interactions should predominantly govern the energy transfer mechanism among them, and the critical distance between them is ~31 Å.  相似文献   

5.
《Surface science》2003,470(1-2):L840-L846
Chemisorption of a family of six chloroethylenes (C2H3Cl, 1,1-C2H2Cl2, cis-1,2-C2H2Cl2, trans-1,2-C2H2Cl2, C2HCl3, and C2Cl4) on Si(1 1 1)7 × 7 at room temperature (RT) has been investigated by vibrational electron energy loss spectroscopy (EELS). The characteristic vibrational EELS features have been used to identify the prominent surface species upon RT adsorption. Like ethylene, C2H3Cl has been found to predominantly adsorb in a di-σ bonding geometry to the Si surface, while 1,1-C2H2Cl2, cis- and trans-1,2-C2H2Cl2, C2HCl3 and, to a lesser extent, C2Cl4 appear to undergo dechlorination upon adsorption to form chlorinated vinyl adspecies involving single-σ bonding structures. Evidence of vinylidene (>CCH2) has been obtained for the first time on a semiconductor surface for the adsorption of 1,1-C2H2Cl2. The present work illustrates that the molecular structure and the Cl content of chloroethylenes play a crucial role in controlling not only the adsorption geometry but also the extent of dechlorination and the resulting adspecies upon RT adsorption on Si(1 1 1).  相似文献   

6.
Magnetic and structural properties of the arrays of 18 nm diameter nanowires of Co and Co90Fe10 electrodeposited in the pores of anodic alumina are investigated. Arrays of Co and Co90Fe10 nanowires show perpendicular magnetic anisotropy and textured crystallographic behaviour. Coercivity Hc (⊥) and remanence Mr/Ms (⊥) values of 2275 Oe (Co90Fe10); 1188 Oe (Co) and 96% (Co90Fe10), 81% (Co) are observed. The continuous films of Co and Co90Fe10 on Cu substrates show in plane magnetic anisotropy and coercivity values between 109 and 288 Oe.  相似文献   

7.
A facile solvothermal method is developed for synthesizing layered Co–Ni hydroxide hierarchical structures by using hexamethylenetetramine (HMT) as alkaline reagent. The electrochemical measurements reveal that the specific capacitances of layered bimetallic (Co–Ni) hydroxides are generally superior to those of layered monometallic (Co, Ni) hydroxides. The as-prepared Co0.5Ni0.5 hydroxide hierarchical structures possesses the highest specific capacitance of 1767 F g−1 at a galvanic current density of 1 A g−1 and an outstanding specific capacitance retention of 87% after 1000 cycles. In comparison with the dispersed nanosheets of Co–Ni hydroxide, layered hydroxide hierarchical structures show much superior electrochemical performance. This study provides a promising method to construct hierarchical structures with controllable transition-metal compositions for enhancing the electrochemical performance in hybrid supercapacitors.  相似文献   

8.
The transition metal-doped spinel cathode materials, LiM0.5Mn1.5O4 (M=Ni. Co, Cr) were prepared by solid-state reaction. The structure and morphology of the samples were investigated by X-ray diffraction, Rietveld refinement and scanning electron microscopy (SEM). The diffraction peaks of all the samples corresponded to a single phase of cubic spinel structure with a space group Fd3m. Field-emission SEM shows octahedron like shapes and the primary particles size was between 500 nm and 2 μm. Oxidation states of Ni, Co and Cr were found to be 2+, 2+ and 3+ as revealed by X-ray photoelectron spectroscopy. During discharging, LiNi0.5Mn1.5O4 and LiCo0.5Mn1.5O4 sample shows more than 130 mAh/g between 3.5 and 5.2 V at a current density of 0.65 mA/cm2 and well developed plateau around 5 V, respectively.  相似文献   

9.
The influence of chlorinated paraffin/titanium (C24H29Cl21/Ti) additives on burning and radiance performances of Magnesium/Teflon/Viton™ (MTV) foil-type was investigated via a high-speed camera, high-temperature differential thermobalance, far-infrared thermal imager and Fourier Transform Infrared (FTIR) remote-sensing spectrometer. We found that the burning temperature, radiance brightness, radiance area and radiance intensity after addition of C24H29Cl21/Ti are improved by 124–196 °C (8–13%), 300–475 W·m−2·sr−1 (12–19%), 943–1422 mm2 (67–101%) and 3.17–4.99 W·sr−1 (88–138%), respectively, and are maximized at the addition ratio of 10%. The substances formed by adding C24H29Cl21/Ti could improve the middle and far infrared radiation.  相似文献   

10.
Exploiting the mechanically controllable break junction technique, we have measured the conductance of atom-sized contacts of Fe, Co, and Ni at room temperature under ultrahigh vacuum conditions. The conductance histogram of Fe exhibits a broad peak around 2.5 G0 (G0  2e2/h), whereas those of Co and Ni show no conductance peaks. However, the histograms of Co and Ni display different structures: While the Co histogram is simply flat, the Ni histogram reveals an appreciable background. Our experimental results are compared with previous results obtained at cryogenic and room temperatures, and the observed peak missing in our room-temperature histograms of Co and Ni is discussed.  相似文献   

11.
《Solid State Ionics》2009,180(40):1646-1651
NiO–C nanocomposite was prepared by a spray pyrolysis method using a mixture of Ni(NO3)2 and citric acid solution at 600 °C. The microstructure and morphology of the NiO–C composite were characterized by means of X-ray diffraction (XRD), transmission electron microscopy (TEM), energy dispersive spectroscopy (EDS) mapping, and thermogravimetric analysis (TGA). The results showed that the NiO nanoparticles were surrounded by amorphous carbon. Electrochemical tests demonstrated that the NiO–C nanocomposites exhibited better capacity retention (382 mAh g 1 for 50 cycles) than that of pure NiO (141 mAh g 1 for 50 cycles), which was also prepared by spray pyrolysis using only Ni(NO3)2 as precursor. The enhanced capacity retention can be mainly attributed to the NiO–C composite structure, composed of NiO nanoparticles surrounded by carbon, which can accommodate the volume changes during charge–discharge and improve the electrical conductivity between the NiO nanoparticles.  相似文献   

12.
《Solid State Ionics》2006,177(7-8):803-811
The purpose of this study was to synthesize highly dispersed Ni/Al2O3 catalysts and to develop a suitable hydrogen-temperature programmed desorption (H2-TPD) method for the determination of nickel metal surface area, dispersion, and crystallite sizes. Several highly dispersed Ni/Al2O3 catalysts with a Ni loading between 15 and 25 wt.% were synthesized. The reducibility of catalysts was determined by temperature programmed reduction (TPR) experiments. All catalysts exhibited a single reduction peak with a maximum rate of H2 consumption (Tmax in TPR) occurring below 450 °C. Three different H2-TPD methods were employed to determine the amount of H2 chemisorbed. In TPD-1, a 10% H2/Ar mixture was used for catalyst pre-reduction and surface saturation by cooling down from Tmax in TPR to room temperature. In TPD-2, the catalyst surface after pre-reduction was flushed with Ar at Tmax in TPR + 10 °C. The TPD-3 was similar to the TPD-2, but used 100% H2 instead of 10% H2/Ar mixture. In all three TPD methods, the profiles exhibited 2 domains of H2 desorption peaks, one below 450 °C, referred to as type-1 peaks, and attributed to H2 desorbed from exposed fraction of Ni atoms, and the other above 450 °C, denoted as type-2 peaks, and assigned to the desorption of H2 located in the subsurface layers and/or to spillover H2. Flushing the reduced catalyst surface in Ar at Tmax in TPR + 10 °C in TPD-2 and TPD-3 removed most of the H2 located in the subsurface layers/ spillover H2. The amount of H2 chemisorbed to form a monolayer on the reduced Ni/Al2O3 catalysts was determined quantitatively from the TPD peak areas of type-1 peaks in TPD-1, and from both type-1 and type-2 peaks in TPD-2 and TPD-3. The Ni metal surface area, dispersions and crystallite sizes were calculated from the chemisorption data and the values were compared with those obtained using the static chemisorption method. Both TPD-2 and TPD-3 gave chemisorption results similar to that obtained from the static method.  相似文献   

13.
Hydrogen peroxide (H2O2) and hydroperoxy (HO2) reactions present in the H2O2 thermal decomposition system are important in combustion kinetics. H2O2 thermal decomposition has been studied behind reflected shock waves using H2O and OH diagnostics in previous studies (Hong et al. (2009) [9] and Hong et al. (2010) [6,8]) to determine the rate constants of two major reactions: H2O2 + M  2OH + M (k1) and OH + H2O2  H2O + HO2 (k2). With the addition of a third diagnostic for HO2 at 227 nm, the H2O2 thermal decomposition system can be comprehensively characterized for the first time. Specifically, the rate constants of two remaining major reactions in the system, OH + HO2  H2O + O2 (k3) and HO2 + HO2  H2O2 + O2 (k4) can be determined with high-fidelity.No strong temperature dependency was found between 1072 and 1283 K for the rate constant of OH + HO2  H2O + O2, which can be expressed by the combination of two Arrhenius forms: k3 = 7.0 × 1012 exp(550/T) + 4.5 × 1014 exp(?5500/T) [cm3 mol?1 s?1]. The rate constants of reaction HO2 + HO2  H2O2 + O2 determined agree very well with those reported by Kappel et al. (2002) [5]; the recommendation therefore remains unchanged: k4 = 1.0 × 1014 exp(?5556/T) + 1.9 × 1011+exp(709/T) [cm3 mol?1 s?1]. All the tests were performed near 1.7 atm.  相似文献   

14.
《Solid State Ionics》2006,177(26-32):2407-2411
Electrical conduction of Sr-doped LaP3O9 ([Sr]/{[La] + [Sr]} = 2–10 mol%) was investigated under 0.4–5 kPa of p(H2O) and 0.01–100 kPa of p(O2) or 0.3–3 kPa of p(H2) at 573–973 K. Sr-doped LaP3O9 showed apparent H/D isotope effect on conductivity regardless of the Sr-doping level under both H2O/O2 oxidizing and H2/H2O reducing conditions at investigated temperatures. Conductivities of the material were almost independent of p(O2) and p(H2O). These results demonstrated that the Sr-doped LaP3O9 exhibited protonic conduction under wide ranges of p(O2), p(H2O) and temperature. The conductivity of the Sr-doped LaP3O9 increased with increasing Sr concentration up to its solubility limit, ca. 3 mol%, while the further Sr-doping slightly degraded the conductivity. These indicate that Sr2+ substitution for La3+ leads to proton dissolution into the material and induced protonic conduction. Conductivities of the 3 mol% Sr-doped sample were 2 × 10- 6–5 × 10 4 S cm 1 at 573–973 K.  相似文献   

15.
Cobalt doping between 2 and 10 at.% was utilized to lower the required sintering temperature of materials in the series BaCe0.5Zr0.4(Y,Yb)0.1 ? yCoyO3 ? δ to between 1373 and 1698 K. The required sintering temperature decreased with increasing Co content; however, significant electronic conductivity was observed in both oxidizing and reducing environments for materials with 10 at.% Co. This was accompanied by a loss of chemical stability in H2O/H2 and CO2 environments. BaCe0.5Zr0.4Yb0.07Co0.03O3 ? δ was stable in these environments and provided the highest proton conductivity of the materials tested, 1.98 × 10? 3 S/cm at 923 K in humidified H2. Measurements in a hydrogen concentration cell indicated that the total ionic transference number for this material was between 0.86 and 1.00 with proton transference number between 0.84 and 0.75 at 773 and 973 K respectively. Under oxidizing conditions, the ionic transference number decreased to below 0.10. The grain boundary resistance dominated the total conductivity at low temperatures but was found to decrease with increased sintering temperature due to grain growth.  相似文献   

16.
《Solid State Ionics》2006,177(26-32):2639-2642
We introduce a newly developed combinatorial electrostatic atomization system, “M-ist Combi,” and demonstrate the effectiveness of the system by establishing a pseudo-ternary Li–Ni–Co oxide phase diagram. After heating the starting materials with compositions in the range of 0.4  Li / (Li + Ni + Co)  0.6 at 973 K for 3 h, the diffraction of all of the products was indexed as single-phase with layer-type hexagonal structures such as LiCoO2 and LiNiO2. As the substitution quantity of Co to the Ni site increased, the value of 2θ shifted to a high-angle. By combining the M-ist Combi system with combinatorial XRD apparatus, we successfully completed the high-throughput sample preparation and phase identification of over 150 samples in one day.  相似文献   

17.
《Current Applied Physics》2010,10(2):655-658
We have quantitatively investigated the Hall effect in [Co, CoFe/Pt] multilayer films. The [Co, CoFe/Pt] multilayers exhibit large spontaneous Hall resistivity (ρH) and Hall angle (ρH/ρ). Even though the Hall resistivity in [Co, CoFe/Pt] multilayer films (2.7–4 × 10−7 Ω cm) is smaller than that of amorphous RE–TM alloy films which show large spontaneous Hall resistivity (<2 × 10−6 Ω cm), the Hall angle of multilayer (6–8%) is almost twice than that in amorphous rare earth–transition metal alloy films (∼3%). The Hall angle provides evidence of the effects of the exchange interaction of the Hall scattering. The exchange is between conduction electron spins and the localized spins of the transition metal. The large Hall angle of [Co, CoFe/Pt] multilayer can be considered due to the high spin polarization and high Curie temperature of Co and CoFe transition metal layers. Even though the role of interfaces and surfaces in the magnetic properties of multilayer films may dominate that of the bulk, the Hall effects in [Co, CoFe/Pt] multilayer may be mainly dominated by the bulk effect.  相似文献   

18.
Polycrystalline single Co nanowires are prepared by electron beam lithography on GaAs substrates at room temperature. The width of the Co nanowires is varied between 150 and 4000 nm. Magnetoresistance measurements are carried out in a temperature range between 1.5 and 45 K applying magnetic fields μ0H up to 4.5 T parallel and perpendicular to the current direction. The in plane (longitudinal) magnetoresistance (MR) shows pronounced features at magnetic fields Hc (coercive fields) indicating the magnetization reversal process. From the MR-curves we determined Hc as a function of the angle α between current and field direction (from in plane to out of plane) and of the width w of the Co nanowires. The Hc=Hc(α,w) behavior allows to discuss the reversal process in more detail.  相似文献   

19.
Ni-containing anode is currently used with many electrolytes of solid oxide fuel cells (SOFCs). However, Ni is easily oxidized and deteriorates the LaGaO3-based electrolyte. A La-doped SrTiO3 (LST, La0.2Sr0.8TiO3) is a candidate as an anode material to solve the Ni poisoning problem in LaGaO3-based SOFC. In this study, a single-phase LST and an LST-Gd0.2Ce0.8O2 ? δ (GDC) composite were tested as the possible anodes on La0.9Sr0.1Ga0.8Mg0.2O3 ? δ (LSGM) electrolyte. In order to further improve the anodic performance, Ni was impregnated into the LST-GDC composite anode. The performance was examined from 600 °C to 800 °C by measuring impedance of the electrolyte-supported, symmetric (anode/electrolyte/anode) cells. A polarization resistance (Rp) of LST-GDC anode was much reduced from that of LST anode. When Ni was impregnated into LST-GDC composite, the Rp value was further reduced to ~ 10% of the single-phase LST anode, and it was 1 Ωcm2 at 800 °C in 97% H2 + 3% H2O atmosphere. A single cell with Ni-impregnated LST-GDC as an anode, Ba0.5Sr0.5Co0.8Fe0.2O3 ? δ (BSCF) as a cathode and LSGM as an electrolyte exhibited the maximum power density of 275 mW/cm2 at 800 °C, increased from ~ 60 mW/cm2 for the cell using the LST-GDC as an anode. Thus, LST-GDC composite is promising as a component of anode.  相似文献   

20.
IV curves showing negative differential resistance (NDR) are reported for single crystals of Co2FeO2BO3 at 315 K and 290 K and for Fe3O2BO3 at 300 K, 260 K and 220 K. Resistivity measurements are presented for both systems, parallel and perpendicular to the c axis, in the range 315–120 K. The high hysteretic behavior of the IV curves in Co2FeO2BO3 around room temperature is discussed and the heat dissipated is estimated, suggesting an increase in the sample temperature of almost 22 K for the IV curve at 315 K and a dominant contribution of Joule self-heating for the observed NDR. In contrast, insignificant hysteresis is observed on the IV curves of Fe3O2BO3 around room temperature. The depinning of charge order domains is suggested as the main contribution to the NDR phenomenon for Fe3O2BO3. The high reproducibility of the NDR in the Fe3O2BO3 single crystal allows its use as a low frequency oscillator, as it is demonstrated.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号